Next Article in Journal
Recent Applications of Pd-Catalyzed Suzuki–Miyaura and Buchwald–Hartwig Couplings in Pharmaceutical Process Chemistry
Next Article in Special Issue
13C NMR Spectroscopic Studies of Intra- and Intermolecular Interactions of Amino Acid Derivatives and Peptide Derivatives in Solutions
Previous Article in Journal
1-(4-Nitrophenyl)-1H-1,2,3-Triazole-4-carbaldehyde: Scalable Synthesis and Its Use in the Preparation of 1-Alkyl-4-Formyl-1,2,3-triazoles
Previous Article in Special Issue
Synthesis of 1-Trifluorometylindanes and Close Structures: A Mini Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Micellar Suzuki Cross-Coupling between Thiophene and Aniline in Water and under Air

1
Department of Materials Science, University of Milano-Bicocca, Via R. Cozzi 55, 20125 Milano, Italy
2
Renewable Energy, Magnetic Fusion and Material Science Research Center, Istituto Donegani, Eni S.p.A., Via Fauser 4, 28100 Novara, Italy
*
Author to whom correspondence should be addressed.
Organics 2021, 2(4), 415-423; https://doi.org/10.3390/org2040025
Submission received: 3 November 2021 / Revised: 8 December 2021 / Accepted: 10 December 2021 / Published: 16 December 2021
(This article belongs to the Special Issue Feature Papers in Organics)

Abstract

:
The Suzuki–Miyaura cross-coupling reaction plays a fundamental role in modern synthetic organic chemistry, both in academia and industry. For this reason, scientists continue to search for new, more effective, cheaper and environmentally friendly procedures. Recently, micellar synthetic chemistry has been demonstrated to be an excellent strategy for achieving chemical transformations in a more efficient way, thanks to the creation of nanoreactors in aqueous environments using selected surfactants. In particular, the cheap and commercially available surfactant Kolliphor EL (a polyethoxylated castor oil derivative) has been used with success to achieve metal-catalyzed transformations in water with high yields and short reaction times, with the advantage of using air-sensitive catalysts without the need for inert atmosphere. In this work, the Kolliphor EL methodology was applied to the Suzuki cross-coupling reaction between thiophene and aniline, using the highly effective catalyst Pd(dtbpf)Cl2. The cross-coupling products were achieved at up to 98% yield, with reaction times of up to only 15 min, working at room temperature and without the need for inert atmosphere.

Graphical Abstract

1. Introduction

Since its discovery in 1979 [1], the Suzuki–Miyaura cross-coupling reaction has been used as an election tool for organic synthesis, both in discovery chemistry and manufacturing processes, due to its broad functional-group tolerance; its reproducibility; and its use of stable, environmentally benign, inexpensive and readily prepared boron reagents [2,3,4,5,6]. Despite the huge advances attained after its introduction, there is still room for improvement, in particular regarding the use of ecocompatible reaction media and shorter reaction times to minimize energy consumption in scaling processes. Recently, surfactant-mediated micellar catalysis has emerged in the field of organic chemistry, allowing synthetic transformation in water with excellent results in term of reaction kinetics, yields and ecocompatibility [7,8,9,10]. The key is the use of selected surfactants able to form association colloids (micelles) of organic molecules in water that behave similarly to nano- and microreactors, whereby organic transformations can occur efficiently due to the mutual proximity of the reagent and catalyst, somehow mimicking what nature has done within cells for billions of years. Thus, C-C and C-N bond-forming reactions, and several other organic transformations, have been successfully achieved with micellar procedures, with substantial improvements with respect to classic organic solvent methodologies [11,12,13,14,15,16]. Within these, the commercially available surfactant Kolliphor EL (a polyethoxylated castor oil, shown in Figure 1) has been proven to form oxygen free micelles, enabling cross-coupling reactions without the need for deoxygenation of the reaction environment [17,18].
Within this context, we were interested in the preparation of thiophene-substituted anilines to be used as precursors for the synthesis of complex heterocyclic systems [19,20,21,22]. Beyond this, thiophene-substituted anilines also serve to produce conductive polymers [23] and as ligand precursors for coordination chemistry [24,25,26]. The obvious strategy for their production involves the cross-coupling processes starting from selected aniline and thiophene precursors, with the Suzuki–Miyaura process being the reaction of choice. In this work, we explored the Kolliphor EL methodology for the Suzuki cross-coupling reaction between thiophenes and anilines, using the highly effective catalyst Pd(dtbpf)Cl2 [27], in comparison to reported classic procedures. The reactions of 2-, 3- and 4-bromoaniline with 2- and 3-thienyl boronic acids and of 2- and 3-bromothiophene with 2-, 3- and 4-aniline boron reagents were tested, as well as 2,4-, 3,4-, 2,5- and 3,5-dibromoaniline with 2- and 3-thienyl boronic acids. Excellent results were obtained in terms of the isolated yields and reaction kinetics with respect to classical organic solvent procedures.

2. Materials and Methods

2.1. General Information

All reagents and solvents were purchased from commercial sources (Merck Life Science S.r.l., Milan, Italy, Fluorochem Ltd., Hadfield, United Kingdom, and TCI Europe N.V., Zwijndrecht, Belgium) and used without further purification. Ultra-Turrax T25 (IKA-Werke GmbH and Co. KG, Staufen, Germany) was used for the Kolliphor EL–toluene mixing. NMR spectra (copies in the Supplementary Materials) were recorded with a Bruker AVANCE III HD 400 MHz spectrometer (Bruker corp., Billerica, MA, USA) (1H:400 MHz, 13C:101 MHz). Chemical shifts (δ) are expressed in parts per million (ppm). Splitting patterns are indicated as follows: s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad. IR spectra were recorded with a Perkin Elmer Spectrum 100 FT-IR spectrometer equipped with a universal ATR sampling accessory (PerkinElmer Inc., Waltham, MA, USA). Melting points were measured with a Stanford Research Systems Optimelt apparatus (SRS, Sunnyvale, CA, USA). Elemental analyses were obtained with an Elementar vario MICRO cube instrument (Elementar Analysensysteme GmbH, Langenselbold, Germany).

2.2. General Procedure 1: Micellar Suzuki Cross-Coupling between Monobromoanilines and Thiophene Boronic Acids

A mixture of bromoaniline 1a–c (0.5 mmol), thiophene boronic acid 2a–b (0.6 mmol), Pd(dtbpf)Cl2 (0.01 mmol), Et3N (1 mmol) and aqueous Kolliphor EL (2 mL, 1.97% H2O) was stirred (500 rpm) at r.t. for the time stated in the text. EtOH was then added (approximately 10 mL, till the reaction mixture was homogeneous) and the solvents were removed under reduced pressure. The residue was purified by flash column chromatography (SiO2, CH2Cl2/n-hexane 8:2) to afford the desired product 3aa-cb in pure form.
2-(2-Thienyl)aniline (3aa). Yellowish solid; mp 35 °C; Rf = 0.28; IR (ATR): 3445, 3363, 3101, 3069, 3021 1612, 1488, 1453, 1302, 1245, 1195, 1158, 1142, 1036, 957, 849, 818, 747, 695, 617, 543 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.38 (dd, J = 5.1, 0.7 Hz, 1H), 7.34 (dd, J = 7.6, 1.1 Hz, 1H), 7.27–7.14 (m, 3H), 6.85 (t, J = 7.5 Hz, 1H), 6.80 (d, J = 8.0 Hz, 1H), 4.03 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 144.1, 141.2, 131.0, 129.1, 127.6, 125.9, 125.3, 120.0, 118.6, 115.9; The spectroscopic data are in accordance to those reported in the literature [25].
2-(3-Thienyl)aniline (3ab). Light brown solid; mp 40 °C; Rf = 0.26; IR (ATR) 3445, 3356, 3098, 3073, 3028, 1613, 1487, 1452, 1360, 1298, 1262, 1192, 1157, 1081, 1048, 1022, 937, 896, 861, 836, 786, 746, 701, 650, 550 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.44 (dd, J = 4.9, 3.0 Hz, 1H), 7.39 (dd, J = 2.8, 1.3 Hz, 1H), 7.29 (dd, J = 4.8, 1.1 Hz, 1H), 7.23 (dd, J = 7.6, 1.5 Hz, 1H), 7.16 (td, J = 7.7, 1.5 Hz, 1H), 6.83 (t, J = 7.5 Hz, 1H), 6.78 (dd, J = 8.0, 0.9 Hz, 1H), 3.85 (br, 2H). The spectroscopic data are in accordance to those reported in the literature [26].
3-(2-Thienyl)aniline (3ba). Yellowish oil; Rf = 0.27; IR (ATR): 3438, 3356, 3101, 3069, 3040, 1617, 1598, 1581, 1485, 1453, 1304, 1230, 1198, 1168, 1079, 1048, 993, 858, 827, 775, 688, 609, 558 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.16 (dd, J = 3.6, 1.1 Hz, 1H), 7.14 (dd, J = 5.1, 1.1 Hz, 1H), 7.05 (t, J = 7.8 Hz, 1H), 6.97–6.90 (m, 2H), 6.81 (t, J = 2.0 Hz, 1H), 6.48 (ddd, J = 7.9, 2.3, 0.9 Hz, 1H), 3.47 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 146.9, 144.7, 135.4, 129.9, 127.9, 124.6, 123.1, 116.5, 114.4, 112.6; Anal. Calcd. for C10H9NS: C, 68.53; H, 5.18; N, 7.99. Found: C, 68.58; H, 5.16; N, 7.95.
3-(3-Thienyl)aniline (3bb). White solid; mp 85 °C; Rf = 0.26; IR (ATR): 3398, 3300, 3204, 3099, 2922, 2852, 1597, 1583, 1530, 1494, 1460, 1367, 1286, 1227, 1192, 1091, 993, 864, 843, 765, 681, 646, 560, 533 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.32 (dd, J = 2.4, 1.9 Hz, 1H), 7.28–7.24 (m, 2H), 7.10 (t, J = 7.8 Hz, 1H), 6.92 (ddd, J = 7.6, 1.6, 1.0 Hz, 1H), 6.84–6.82 (m, 1H), 6.54 (ddd, J = 7.9, 2.3, 0.9 Hz, 1H), 3.58 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 146.8, 142.5, 136.9, 129.7, 126.4, 126.0, 120.2, 117.1, 114.0, 113.2; Anal. Calcd. for C10H9NS: C, 68.53; H, 5.18; N, 7.99. Found: C, 68.55; H, 5.14; N, 8.01.
4-(2-Thienyl)aniline (3ca). Light brown solid; mp 76 °C; Rf = 0.28; IR (ATR): 3440, 3356, 3195, 3097, 2922, 2852, 1614, 1603, 1531, 1499, 1429, 1407, 1286, 1256, 1182, 1132, 1081, 1050, 954, 846, 809, 698, 660, 636, 571 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.43 (d, J = 8.4 Hz, 1H), 7.17 (t, J = 4.4 Hz, 1H), 7.04 (dd, J = 4.7, 3.9 Hz, 1H), 6.69 (d, J = 8.4 Hz, 1H), 3.73 (s, 1H). The spectroscopic data are in accordance to those reported in the literature [28].
4-(3-Thienyl)aniline (3cb). Yellow solid; mp 97 °C; Rf = 0.26; IR (ATR): 3401, 3306, 3207, 3096, 3039, 1625, 1603, 1537, 1504, 1439, 1365, 1268, 1256, 1204, 1189,1130, 1094, 855, 827, 775, 687, 671, 623, 567, 509 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.45–7.41 (m, 2H), 7.38–7.33 (m, 2H), 7.31 (dd, J = 2.6, 1.6 Hz, 1H), 6.75–6.70 (m, 2H), 3.70 (br, 2H). The spectroscopic data are in accordance to those reported in the literature [24].

2.3. General Procedure 2: Micellar Suzuki Cross-Coupling between Anilines Boronic Acids and Esters and Bromo-Thiophenes

A mixture of aniline boronic acid 4a–b or ester 4c (0.6 mmol), bromo-thiophene 5a–b (0.5 mmol), Pd(dtbpf)Cl2 (0.01 mmol), Et3N (1 mmol) and Kolliphor EL (1.97%, H2O)/toluene 9:1 (2 mL, premixed with Ultra-Turrax at 20,000 rpm for 5 min), was stirred (500 rpm) at 60 °C for the time stated in the text. EtOH was then added (approximately 10 mL, until the reaction mixture was homogeneous) and the solvents were removed under reduced pressure. The residue was purified by flash column chromatography (SiO2, CH2Cl2/n-hexane8:2) to afford the desired product 3aa–cb in pure form.

2.4. General Procedure 3: Micellar Suzuki Cross-Coupling between Dibromoanilines and Thiophene Boronic Acids

A mixture of dibromoaniline 6a–d (0.5 mmol), thiophene boronic acid 2a–b (1.2 mmol), Pd(dtbpf)Cl2 (0.01 mmol), Et3N (2 mmol) and Kolliphor EL (1.97%, H2O)/toluene 9:1 (2 mL, premixed with Ultra-Turrax at 20,000 rpm for 5 min) was stirred (500 rpm) at 60 °C for the time stated in the text. EtOH was then added (approximately 10 mL, until the reaction mixture was homogeneous) and the solvents were removed under reduced pressure. The residue was purified by flash column chromatography (SiO2, CH2Cl2/n-hexane 8:2) to afford the desired product 7aa–db in pure form.
2,4-di-(2-Thienyl)aniline (7aa). Yellowish oil; Rf = 0.34; IR (ATR): 3459, 3370, 3208, 3101, 3068, 3021, 1615, 1534, 1491, 1429, 1405, 1343, 1295, 1260, 1233, 1212, 1183, 1157, 1078, 1048, 944, 888, 845, 811, 689, 616, 564 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.54 (d, J = 2.2 Hz, 1H), 7.42 (dd, J = 8.3, 2.2 Hz, 1H), 7.39 (dd, J = 5.2, 1.1 Hz, 1H), 7.24 (dd, J = 3.5, 1.1 Hz, 1H), 7.21–7.17 (m, 2H), 7.15 (dd, J = 5.1, 3.5 Hz, 1H), 7.05 (dd, J = 4.8, 3.9 Hz, 1H), 6.78 (d, J = 8.3 Hz, 1H), 4.09 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 144.5, 143.7, 140.5, 128.7, 127.9, 127.6, 126.9, 126.2, 125.6, 125.2, 123.4, 121.6, 120.2, 116.1; Anal. Calcd. For C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.26; H, 4.36; N, 5.51.
2,5-di-(2-Thienyl)aniline (7ba). Yellowish solid; mp 80 °C; Rf = 0.36; IR (ATR): 3463, 3370, 3094, 3064, 2923, 2852, 1614, 1599, 1484, 1427, 1413, 1353, 1318, 1291, 1260, 1231, 1193, 1128, 1079, 1041, 940, 870, 844, 834, 814, 713, 690, 572 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.37 (dd, J = 5.1, 1.0 Hz, 1H), 7.34–7.30 (m, 2H), 7.29 (dd, J = 5.1, 1.0 Hz, 1H), 7.24 (dd, J = 3.5, 1.0 Hz, 1H), 7.15 (dd, J = 5.1, 3.5 Hz, 1H), 7.11–7.06 (m, 2H), 7.03 (d, J = 1.8 Hz, 1H), 4.10 (br, 2H); The spectroscopic data are in accordance to those reported in the literature [24].
3,4-di-(2-Thienyl)aniline (7ca). Light brown solid; mp 75 °C; Rf = 0.32; IR (ATR): 3464, 3369, 3189, 3094, 3064, 3030, 1610, 1598, 1483, 1449, 1427, 1413, 1353, 1318, 1292, 1260, 1232, 1198, 1079, 845, 834, 813, 714, 690, 571, cm−1; 1H NMR (400 MHz, CDCl3) δ 7.21 (d, J = 8.3 Hz, 1H), 7.15 (dd, J = 5.1, 1.2 Hz, 1H), 7.11 (dd, J = 5.1, 1.2 Hz, 1H), 6.87–6.81 (m, 2H), 6.77 (dd, J = 3.5, 1.2 Hz, 1H), 6.73 (d, J = 2.5 Hz, 1H), 6.70 (dd, J = 3.5, 1.2 Hz, 1H), 6.59 (dd, J = 8.2, 2.5 Hz, 1H), 3.68 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 146.2, 143.2, 142.9, 134.7, 132.2, 126.9, 126.8, 126.8, 126.3, 125.7, 125.0, 124.0, 117.1, 114.5; Anal. Calcd. For C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.39; H, 4.30; N, 5.43.
3,5-di-(2-Thienyl)aniline (7da). Yellow solid; mp 140 °C; Rf = 0.35; IR (ATR): 3420, 3307, 3203, 3101, 3066, 2953, 2921, 2851, 1623, 1590, 1525, 1434, 1364, 1319, 1231, 1198, 1078, 1028, 988, 851, 819, 751, 701, 689, 612 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.33 (dd, J = 3.6, 1.1 Hz, 2H), 7.31–7.27 (m, 3H), 7.09 (dd, J = 5.1, 3.6 Hz, 2H), 6.86 (d, J = 1.5 Hz, 2H), 3.78 (br, 2H). The spectroscopic data are in accordance to those reported in the literature [29].
2,4-di-(3-Thienyl)aniline (7ab). Yellow solid; mp 85 °C; Rf = 0.32; IR (ATR): 3443, 3355, 3095, 2954, 2922, 2852, 1615, 1493, 1432, 1400, 1364, 1342, 1294, 1260, 1231, 1200, 1186, 1157, 1084, 928, 857, 840, 828, 777, 737, 686, 633, 573 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.33 (d, J = 2.1 Hz, 1H), 7.29 (dd, J = 4.9, 3.0 Hz, 1H), 7.27–7.23 (m, 2H), 7.21–7.19 (m, 2H), 7.18–7.14 (m, 2H), 6.62 (d, J = 8.2 Hz, 1H), 3.66 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 143.2, 142.3, 139.7, 139.7, 128.4, 128.3, 126.7, 126.3, 126.2, 126.0, 122.8, 122.7, 118.3, 116.1; Anal. Calcd. For C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.28; H, 4.34; N, 5.47.
2,5-di-(3-Thienyl)aniline (7bb). Yellow solid; mp 124 °C; Rf = 0.33; IR (ATR): 3425, 3350, 3097, 2951, 2922, 2852, 1614, 1561, 1493, 1433, 1357, 1314, 1294, 1267, 1196, 1183, 1084, 947, 862, 846, 778, 738, 705, 649, 609, 583, 552, 520 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.46–7.42 (m, 2H), 7.41 (dd, J = 3.0, 1.3 Hz, 1H), 7.39–7.36 (m, 1H), 7.29 (dd, J = 4.9, 1.3 Hz, 1H), 7.25 (d, J = 7.8 Hz, 1H), 7.05 (dd, J = 7.8, 1.8 Hz, 1H), 7.01 (d, J = 1.7 Hz, 1H), 3.92 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 144.1, 142.2, 139.5, 136.1, 130.6, 128.3, 126.4, 126.1, 126.0, 122.5, 121.5, 120.2, 117.0, 113.6; Anal. Calcd. For C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.40; H, 4.31; N, 5.42.
3,4-di-(3-Thienyl)aniline (7cb). Brown solid; mp 79 °C; Rf = 0.29; IR (ATR): 3442, 3356, 3096, 2955, 2920, 2850, 1602, 1576, 1537, 1476, 1359, 1303, 1261, 1239, 1189, 1080, 1024, 940, 855, 845, 825, 779, 737, 693, 679, 662, 648, 612, 594, 566, 518 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.16 (d, J = 8.2 Hz, 1H), 7.08–7.01 (m, 2H), 6.95 (dd, J = 3.0, 1.3 Hz, 1H), 6.85 (dd, J = 3.0, 1.3 Hz, 1H), 6.69 (dd, J = 5.0, 1.2 Hz, 1H), 6.67 (d, J = 2.5 Hz, 1H), 6.64 (dd, J = 5.0, 1.3 Hz, 1H), 6.58 (dd, J = 8.2, 2.5 Hz, 1H), 3.55 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 144.7, 141.2, 141.1, 135.2, 130.2, 128.1, 127.9, 124.9, 123.5, 123.2, 121.6, 120.6, 115.5, 113.2; Anal. Calcd. For C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.44; H, 4.28; N, 5.39.
3,5-di-(3-Thienyl)aniline (7db). White solid; mp 155 °C; Rf = 0.32; IR (ATR): 3424, 3302, 3203, 3104, 3090, 2953, 2923, 2853, 1625, 1593, 1526, 1449, 1374, 1352, 1314, 1274, 1226, 1189, 1093, 992, 933, 875, 848, 834, 773, 684, 643, 619, 574 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.38–7.34 (m, 2H), 7.32–7.27 (m, 4H), 7.14 (t, J = 1.5 Hz, 1H), 6.77 (d, J = 1.5 Hz, 2H), 3.66 (br, 2H); 13C NMR (101 MHz, CDCl3) δ 147.1, 142.4, 137.5, 126.5, 126.1, 120.5, 115.7, 112.2; Anal. Calcd. for C14H11NS2: C, 65.33; H, 4.31; N, 5.44. Found: C, 65.28; H, 4.34; N, 5.47.

3. Results and Discussion

We started our investigation by reacting 2-bromoaniline 1a with 2-thienylboronic acid 2a (Scheme 1), taking a cue from the optimized conditions reported earlier [18]. Thus, 1a (0.5 mmol), 2a (0.6 mmol), Pd(dtbpf)Cl2 (2% mol) and Et3N (1 mmol) were added to 2 mL of a 2 w% Kolliphor EL solution in water and stirred at rt under air. Surprisingly, after only 15 min, TLC control showed no presence of starting materials. Proceeding with the reaction workup, 3aa was isolated at 86% chemical yield (Table 1, Entry 1). Under these conditions, thienyl-anilines 3ab, 3bb, 3ca and 3cb (Scheme 1) were obtained at yields of 81%, 88%, 91% and 94%, respectively (Table 1, Entries 2, 4, 5 and 6, respectively). The reaction of 3-bromoaniuline 1b with 2-thienyl boronic acid 2a was less effective (Table 1, Entry 3); extending the reaction time from 15 to 60 min was enough to increase the yield from 64% to 96% (Table 1, Entry 7). For comparison, we reported in Table 1 (Entries 8–10) classical organic solvent procedures present in the literature carried out with the same class of Pd catalyst, i.e., L2PdCl2 [23,25,30]. The yields were comparable, even if in two cases they were slightly lower, although classic procedures require at least 12 h at higher temperature to complete the transformations and under inert atmosphere.
Stimulated by these results, we tested our methodology inverting the functionalities on starting materials, i.e., boron on aniline and bromine on thiophene (Scheme 2). Surprisingly, the reactions with previous conditions were much less effective (Table 2, Entries 1–4), as starting materials consuming needed 20 h and products were isolated in poor yields (22–45%). In the past, we overcame this unfavorable behavior by resorting to the cosolvent approach [19], i.e., the addition of a 10 v% of water immiscible organic solvent leading to the formation of a micro-emulsion, whereby the cores of the micelles swell, giving more space for the reactions to happen. Thus, we re-ran the reactions by adding a 10 v% of toluene to the Kolliphor EL water solution (vigorous premixing is essential for the micro-emulsion to form). This was beneficial for the kinetics (1 h vs. 20 h), but yields remained low to modest (Table 2, Entries 5–10). Temperature is another factor to consider in micellar procedures, as a slight increment could enormously increase reaction outputs [21]; obviously, the emulsion cloud point must not be exceeded, otherwise the micellar environment will be destroyed [31]. The temperature was increased to 60 °C and products 3aa–cb were obtained in excellent yields, ranging from 88 to 96% (Table 2, Entries 11–16). To prove the combined effect of cosolvent and temperature, we attempted the reaction of 4c with 5b at 60 °C without the addition of toluene, whereby the yield dropped from 96 to 75% (Table 2, Entry 17). Clear advantages of the use of Kolliphor EL methodology respect to reported classical ones in organic solvent and inert atmosphere could be seen also for these kinds of reactions (Table 2, Entries 18 and 19) [32,33].
Finally, the Kolliphor-EL-mediated Suzuki reaction was tested for the thienyl functionalization of di-bromoanilines (Scheme 3). As observed for anilines boronic acids, the use of sole water–surfactant medium at rt resulted in slow kinetics and poor yields. Therefore, we opted for the use of 10 v% toluene as a cosolvent at 60 °C for 1 h (the results are presented in Table 3). These conditions allowed us to obtain di-thienyl-substituted anilines 7aa–db in excellent isolated yields, ranging from 85% for 3,4-di-(3-thienyl)aniline 7cb (Table 3, Entry 7) to 98% for 2,4-di-(2-thienyl)aniline 7aa and 2,4-di-(3-thienyl)aniline 7ab (Table 3, Entries 1 and 5). Importantly, (dtbpf)PdCl2 was used at the same molar ratio of mono-bromoanilines, ending in a 1 mol% ratio with respect to the C-Br bonds present, without loss of efficiency. A comparison with classical procedures was more difficult in this case, as only one procedure is reported in the literature for the sole thienyl functionalization of di-bromoanilines (Table 3, Entry 9) [24]; however, 72 h at 95 °C afforded only 55% yield, confirming once again the advantage of the Kolliphor EL mediated micellar procedure.

4. Conclusions

In summary, we tested the Kolliphor-EL-mediated micellar Suzuki–Miyaura reaction for the production of thienyl-substituted anilines. Mono-thienylanilines were efficiently prepared using a 2 w% Kolliphor EL water solution at rt under air for 15 min starting from bromoanilines, while cosolvent addition, at a slightly higher temperature for 1 h of reaction time, was needed when starting from aniline boronic acid or esters. Under these latter conditions, di-thienylanilines from di-bromoanilines were obtained in the best yields reported to date. This work demonstrated that Suzuki cross-coupling to thienyl substituted anilines can be efficiently performed with the Kolliphor EL micellar system in a very simple and cost-effective way compared to previously reported methods, working at low temperature, with short reaction times and in air.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/org2040025/s1. Images of 1H and 13C NMR spectra for unknown compounds and images of 1H NMR spectra for known compounds.

Author Contributions

Conceptualization, L.V. and A.P.; methodology, L.V.; formal analysis, L.V. and D.Y.; investigation, D.Y., S.T. and E.O.M.; writing—original draft preparation, L.V.; writing—review and editing, A.P., R.P. and P.B.; visualization, L.V. and D.Y.; supervision, A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We gratefully acknowledge Giorgio Patriarca for the NMR analyses and Mauro Monti for laboratory support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miyaura, N.; Yamada, K.; Suzuki, A. A New Stereospecific Cross-Coupling by the Palladium-Catalyzed Reaction of 1-Alkenylboranes with 1-Alkenyl or 1-Alkynyl Halides. Tetrahedron Lett. 1979, 20, 3437–3440. [Google Scholar] [CrossRef] [Green Version]
  2. Neeve, E.C.; Geier, S.J.; Mkhalid, I.A.I.; Westcott, S.A.; Marder, T.B. Diboron(4) Compounds: From Structural Curiosity to Synthetic Workhorse. Chem. Rev. 2016, 116, 9091–9161. [Google Scholar] [CrossRef] [Green Version]
  3. Lennox, A.J.J.; Lloyd-Jones, G.C. Selection of Boron Reagents for Suzuki–Miyaura Coupling. Chem. Soc. Rev. 2014, 43, 412–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef] [Green Version]
  5. Suzuki, A. Recent Advances in the Cross-Coupling Reactions of Organoboron Derivatives with Organic Electrophiles, 1995–1998. J. Organomet. Chem. 1999, 576, 147–168. [Google Scholar] [CrossRef]
  6. Suzuki, A. Organoborane Coupling Reactions (Suzuki Coupling). Proc. Jpn. Acad. Ser. B 2004, 80, 359–371. [Google Scholar] [CrossRef] [Green Version]
  7. Lipshutz, B.H.; Ghorai, S.; Cortes-Clerget, M. The Hydrophobic Effect Applied to Organic Synthesis: Recent Synthetic Chemistry “in Water”. Chem. Eur. J. 2018, 24, 6672–6695. [Google Scholar] [CrossRef]
  8. La Sorella, G.; Strukul, G.; Scarso, A. Recent Advances in Catalysis in Micellar Media. Green Chem. 2015, 17, 644–683. [Google Scholar] [CrossRef]
  9. Dwars, T.; Paetzold, E.; Oehme, G. Reactions in Micellar Systems. Angew. Chem. Int. Ed. 2005, 44, 7174–7199. [Google Scholar] [CrossRef] [PubMed]
  10. Vafakish, B.; Wilson, L.D. A Review on Recent Progress of Glycan-Based Surfactant Micelles as Nanoreactor Systems for Chemical Synthesis Applications. Polysaccharides 2021, 2, 168–186. [Google Scholar] [CrossRef]
  11. Ryu, J.-H.; Jang, C.-J.; Yoo, Y.-S.; Lim, S.-G.; Lee, M. Supramolecular Reactor in an Aqueous Environment:  Aromatic Cross Suzuki Coupling Reaction at Room Temperature. J. Org. Chem. 2005, 70, 8956–8962. [Google Scholar] [CrossRef] [PubMed]
  12. Handa, S.; Wang, Y.; Gallou, F.; Lipshutz, B.H. Sustainable Fe–Ppm Pd Nanoparticle Catalysis of Suzuki-Miyaura Cross-Couplings in Water. Science 2015, 349, 1087–1091. [Google Scholar] [CrossRef] [Green Version]
  13. Ansari, T.N.; Taussat, A.; Clark, A.H.; Nachtegaal, M.; Plummer, S.; Gallou, F.; Handa, S. Insights on Bimetallic Micellar Nanocatalysis for Buchwald–Hartwig Aminations. ACS Catal. 2019, 9, 10389–10397. [Google Scholar] [CrossRef]
  14. Iranpoor, N.; Firouzabadi, H.; Shekarize, M. Micellar Media for the Efficient Ring Opening of Epoxides with CN, N3, NO3, NO2, SCN, Cl and Br Catalyzed with Ce(OTf)4. Org. Biomol. Chem. 2003, 1, 724–727. [Google Scholar] [CrossRef]
  15. Sobhani, S.; Zeraatkar, Z. A New Magnetically Recoverable Heterogeneous Palladium Catalyst for Phosphonation Reactions in Aqueous Micellar Solution. Appl. Organomet. Chem. 2015, 30, 12–19. [Google Scholar] [CrossRef]
  16. Chen, B.-T.; Bukhryakov, K.V.; Sougrat, R.; Rodionov, V.O. Enzyme-Inspired Functional Surfactant for Aerobic Oxidation of Activated Alcohols to Aldehydes in Water. ACS Catal. 2015, 5, 1313–1317. [Google Scholar] [CrossRef]
  17. Mattiello, S.; Monguzzi, A.; Pedrini, J.; Sassi, M.; Villa, C.; Torrente, Y.; Marotta, R.; Meinardi, F.; Beverina, L. Self-Assembled Dual Dye-Doped Nanosized Micelles for High-Contrast Up-Conversion Bioimaging. Adv. Funct. Mater. 2016, 26, 8447–8454. [Google Scholar] [CrossRef]
  18. Mattiello, S.; Rooney, M.; Sanzone, A.; Brazzo, P.; Sassi, M.; Beverina, L. Suzuki–Miyaura Micellar Cross-Coupling in Water, at Room Temperature, and under Aerobic Atmosphere. Org. Lett. 2017, 19, 654–657. [Google Scholar] [CrossRef]
  19. Vaghi, L.; Sanzone, A.; Sassi, M.; Pagani, S.; Papagni, A.; Beverina, L. Synthesis of Fluorinated Acridines via Sequential Micellar Buchwald–Hartwig Amination/Cyclization of Aryl Bromides. Synthesis 2018, 50, 1621–1628. [Google Scholar] [CrossRef]
  20. Vaghi, L.; Coletta, M.; Coghi, P.; Andreosso, I.; Beverina, L.; Ruffo, R.; Papagni, A. Fluorine Substituted Non-Symmetric Phenazines: A New Synthetic Protocol from Polyfluorinated Azobenzenes. Arkivoc 2019, 2019, 340–351. [Google Scholar] [CrossRef] [Green Version]
  21. Yousif, D.; Monti, M.; Papagni, A.; Vaghi, L. Synthesis of Phenazines from Ortho-Bromo Azo Compounds via Sequential Buchwald-Hartwig Amination under Micellar Conditions and Acid Promoted Cyclization. Tetrahedron Lett. 2020, 61, 152511. [Google Scholar] [CrossRef]
  22. Vaghi, L.; Monti, M.; Marelli, M.; Motto, E.; Papagni, A.; Cipolla, L. Photoinduced Porcine Gelatin Cross-Linking by Homobi- and Homotrifunctional Tetrazoles. Gels 2021, 7, 124. [Google Scholar] [CrossRef] [PubMed]
  23. Pardo, M.A.; del Valle, M.A.; Díaz, F.R. Synthesis and Characterization of Aniline and Thiophene and/or Alkylthiophenes Polymers. Polym. Bull. 2015, 72, 2189–2199. [Google Scholar] [CrossRef]
  24. Djukic, B.; Seda, T.; Gorelsky, S.I.; Lough, A.J.; Lemaire, M.T. π-Extended and Six-Coordinate Iron(II) Complexes: Structures, Magnetic Properties, and the Electrochemical Synthesis of a Conducting Iron(II) Metallopolymer. Inorg. Chem. 2011, 50, 7334–7343. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.; Chen, W.; Qiao, Y.; Zhou, G. B2N2-Embedded Polycyclic Aromatic Hydrocarbons with Furan and Thiophene Derivatives Functionalized in Crossed Directions. Chem. Eur. J. 2019, 25, 9326–9338. [Google Scholar] [CrossRef]
  26. Elbert, S.M.; Wagner, P.; Kanagasundaram, T.; Rominger, F.; Mastalerz, M. Boroquinol Complexes with Fused Extended Aromatic Backbones: Synthesis and Optical Properties. Chem. Eur. J. 2016, 23, 935–945. [Google Scholar] [CrossRef] [PubMed]
  27. Colacot, T.J.; Shea, H.A. Cp2Fe(PR2)2PdCl2 (R = i-Pr, t-Bu) Complexes as Air-Stable Catalysts for Challenging Suzuki Coupling Reactions. Org. Lett. 2004, 6, 3731–3734. [Google Scholar] [CrossRef] [PubMed]
  28. Powell, A.B.; Brown, J.R.; Vasudevan, K.V.; Cowley, A.H. Facile Syntheses of Thiophene-Substituted 1,4-Diazabutadiene (α-Diimine) Ligands and Their Conversion to Phosphenium Triiodide Salts. Dalton Trans. 2009, 14, 2521–2527. [Google Scholar] [CrossRef]
  29. Yalçın, E.; Kutlu, Y.C.; Korkmaz, V.; Şahin, E.; Seferoğlu, Z. 2,6-Dicyanoaniline Based Donor-Acceptor Compounds: The Facile Synthesis of Fluorescent 3,5-Diaryl/Hetaryl-2,6-Dicyanoanilines. Arkivoc 2015, 2015, 202–218. [Google Scholar] [CrossRef] [Green Version]
  30. Naveen, K.; Nandakumar, A.; Perumal, P. Synthesis of 4-Alkylidene-Substituted 1,2,3,4-Tetrahydroisoquinolines via Palladium-Catalyzed Carbopalladation/C–H Activation of 2-Bromobenzyl-N-Propargylamines. Synthesis 2015, 47, 1633–1642. [Google Scholar] [CrossRef]
  31. Inoue, T.; Ohmura, H.; Murata, D. Cloud Point Temperature of Polyoxyethylene-Type Nonionic Surfactants and Their Mixtures. J. Colloid Interface Sci. 2003, 258, 374–382. [Google Scholar] [CrossRef]
  32. Li, T.; Yang, Y.; Cheng, C.; Tiwari, A.K.; Sodani, K.; Zhao, Y.; Abraham, I.; Chen, Z.-S. Design, Synthesis and Biological Evaluation of N-Arylphenyl-2,2-Dichloroacetamide Analogues as Anti-Cancer Agents. Bioorg. Med. Chem. Lett. 2012, 22, 7268–7271. [Google Scholar] [CrossRef] [PubMed]
  33. Holzer, M.; Ziegler, S.; Neugebauer, A.; Kronenberger, B.; Klein, C.D.; Hartmann, R.W. Structural Modifications of Salicylates: Inhibitors of Human CD81-Receptor HCV-E2 Interaction. Arch. Pharm. 2008, 341, 478–484. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of Kolliphor EL.
Figure 1. Chemical structure of Kolliphor EL.
Organics 02 00025 g001
Scheme 1. Micellar Suzuki cross-coupling between mono-bromoanilines and thienyl boronic acids.
Scheme 1. Micellar Suzuki cross-coupling between mono-bromoanilines and thienyl boronic acids.
Organics 02 00025 sch001
Scheme 2. Micellar Suzuki cross-coupling between aniline boronic acids and esters and bromo-thiophenes.
Scheme 2. Micellar Suzuki cross-coupling between aniline boronic acids and esters and bromo-thiophenes.
Organics 02 00025 sch002
Scheme 3. Micellar Suzuki cross-coupling between di-bromoanilines and thienyl boronic acids.
Scheme 3. Micellar Suzuki cross-coupling between di-bromoanilines and thienyl boronic acids.
Organics 02 00025 sch003
Table 1. Output of micellar Suzuki cross-coupling between mono-bromoanilines and thienyl boronic acids.
Table 1. Output of micellar Suzuki cross-coupling between mono-bromoanilines and thienyl boronic acids.
EntryBr-AnilineThienyl-B(OH)2TimeTemperatureProductYield% 1
11a2a15 minrt3aa86
21a2b15 minrt3ab81
31b2a15 minrt3ba64
41b2b15 minrt3bb88
51c2a15 minrt3ca91
61c2b15 minrt3cb94
71b2a60 minrt3ba96
8 21a2a12 h70 °C3aa74
9 31a2b24 h80 °C3ab94
10 41c2a16 h80 °C3ca74
1 Isolated yields. 2 Organic solvent procedure with L2PdCl2 catalyst in inert atmosphere reported in [25]. 3 Organic solvent procedure with L2PdCl2 catalyst in inert atmosphere reported in [30]. 4 Organic solvent procedure with L2PdCl2 catalyst in inert atmosphere reported in [23].
Table 2. Output of micellar Suzuki Cross-Coupling between aniline boronic acids and esters and bromo-thiophenes.
Table 2. Output of micellar Suzuki Cross-Coupling between aniline boronic acids and esters and bromo-thiophenes.
EntryAniline-BBr-ThiopheneTimeTemperatureProductYield% 1
1 24a5a20 hrt3aa22
2 24b5a20 hrt3ba34
3 24c5a20 hrt3ca45
4 24a5b20 hrt3ab34
54a5a1 hrt3aa31
64b5a1 hrt3ba40
74c5a1 hrt3ca51
84a5b1 hrt3ab70
94b5b1 hrt3bb50
104c5b1 hrt3cb54
114a5a1 h60 °C3aa95
124b5a1 h60 °C3ba90
134c5a1 h60 °C3ca88
144a5b1 h60 °C3ab95
154b5b1 h60 °C3bb90
164c5b1 h60 °C3cb96
17 24c5b1 h60 °C3cb75
18 34b5a12 h60 °C3ba23
19 44b5b16 h90 °C3bb82
1 Isolated yields. 2 Reaction carried without toluene as cosolvent. 3 Organic solvent procedure in inert atmosphere reported in [32]. 4 Organic solvent procedure in inert atmosphere reported in [33].
Table 3. Output of micellar Suzuki cross-coupling between di-bromoanilines and thienyl boronic acids.
Table 3. Output of micellar Suzuki cross-coupling between di-bromoanilines and thienyl boronic acids.
EntryDiBr-AnilineB-ThiopheneTimeTemperatureProductYield% 1
16a2a1 h60 °C7aa98
26b2a1 h60 °C7ba96
36c2a1 h60 °C7ca90
46d2a1 h60 °C7da90
56a2b1 h60 °C7ab98
66b2b1 h60 °C7bb86
76c2b1 h60 °C7cb85
86d2b1 h60 °C7db89
9 26b2a72 h95 °C7ba55
1 Isolated yields. 2 Organic solvent procedure in inert atmosphere reported in [24].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yousif, D.; Tombolato, S.; Ould Maina, E.; Po, R.; Biagini, P.; Papagni, A.; Vaghi, L. Micellar Suzuki Cross-Coupling between Thiophene and Aniline in Water and under Air. Organics 2021, 2, 415-423. https://doi.org/10.3390/org2040025

AMA Style

Yousif D, Tombolato S, Ould Maina E, Po R, Biagini P, Papagni A, Vaghi L. Micellar Suzuki Cross-Coupling between Thiophene and Aniline in Water and under Air. Organics. 2021; 2(4):415-423. https://doi.org/10.3390/org2040025

Chicago/Turabian Style

Yousif, Dawod, Silvia Tombolato, Elmehdi Ould Maina, Riccardo Po, Paolo Biagini, Antonio Papagni, and Luca Vaghi. 2021. "Micellar Suzuki Cross-Coupling between Thiophene and Aniline in Water and under Air" Organics 2, no. 4: 415-423. https://doi.org/10.3390/org2040025

Article Metrics

Back to TopTop